Electron-Deficient Eu6.5Gd0.5Ge6 Intermetallic: A Layered

Feb 15, 2012 - Department of Chemistry and Chemical Biology, McMaster University, 1280 Main Street West, Hamilton, Ontario L8S 4M1, Canada. Inorg...
0 downloads 0 Views 2MB Size
Article pubs.acs.org/IC

Electron-Deficient Eu6.5Gd0.5Ge6 Intermetallic: A Layered Intergrowth Phase of the Gd5Si4- and FeB-Type Structures Jinlei Yao, Peng Wang, and Yurij Mozharivskyj* Department of Chemistry and Chemical Biology, McMaster University, 1280 Main Street West, Hamilton, Ontario L8S 4M1, Canada S Supporting Information *

ABSTRACT: A novel electron-poor Eu6.5Gd0.5Ge6 compound adopts the Ca7Sn6-type structure (space group Pnma, Z = 4, a = 7.5943(5) Å, b = 22.905(1) Å, c = 8.3610(4) Å, and V = 1454.4(1) Å3). The compound can be seen as an intergrowth of the Gd5Si4-type (Pnma) R5Ge4 (R = rare earth) and FeB-type (Pnma) RGe compounds. The phase analysis suggests that the Eu7−xGdxGe6 series displays a narrow homogneity range of stabilizing the Ca7Sn6 structure at x ≈ 0.5. The structural results illustrate the structural rigidity of the 2∝ [R5X4] slabs (X = p-element) and a possibility for discovering new intermetallics by combining the 2∝ [R5X4] slabs with other symmetry-approximate building blocks. Electronic structure analysis suggests that the stability and composition of Eu6.5Gd0.5Ge6 represents a compromise between the valence electron concentration, bonding, and existence of the neighboring EuGe and (Eu,Gd)5Ge4 phases.



INTRODUCTION Discovery of a giant magnetocaloric effect in Gd5Si2Ge2 stimulates extensive studies on the relationship between the crystal structures and physical properties of the R5X4 family of compounds (where R = rare-earth element and X = p-element).1−11 Except for the tetragonal Zr5Si4- (space group P41212) and hexagonal Ti5Ga4-type (P63/mcm) structures, the R5X4 phases are found to adopt one or more of the eight related structures: Sm5Ge4 (Pnma),12,13 Gd5Si4 (Pnma),12,13 Gd5Si2Ge2 (P21/a),1 Pu5Rh4 (Pnma),5 Eu5As4 (Cmca),11 U2Mo3Si4 (P21/c),9 Gd5Si2Bi2 (I41/acd), and Gd5Si1.5Bi2.5 (P42bc).10 These eight layered structures are constructed from the same building block, the2∝[R5X4] slab, but with various types of interslab X−X bonds (e.g., Gd5Si4 in Figure 1 has all the interslab X−X bonds intact).

interslab X−X bonds can either stretch, break, or reform, thus indicating the structural rigidity of the 2∝ [R5X4] slabs and flexibility of the interslab bonds.2,5,10 These structural features encouraged us to search for new phases beyond the R5X4 family, what would assemble the 2∝ [R5X4] slabs and other symmetry-allowed building blocks into their structures. In this work, we report on the synthesis and characterization of a new series of pseudo-binary Eu7−xGdxGe6 compounds that adopt the Ca7Sn6 structure (Pnma) and display a narrow homogeneity range around the Eu6.5Gd0.5Ge6 composition. The (Eu,Gd)7Ge6 compound, isostructural with Ca7Sn6, shows remarkable structural similarities to the R5X4 ones.



Figure 1. Crystal structure of Gd5Si4-type R5Ge4, Ca7Sn6-type R7Ge6, and FeB-type RGe (R = a mixture of Eu and Gd).

Some R5X4 compounds undergo structural transformations as a function of temperature, magnetic field, or pressure.1,3,4 During such transitions, the 2∝ [R5X4] slabs remain intact but the © 2012 American Chemical Society

EXPERIMENTAL SECTION

Sample Preparation. The Ca7Sn6-type R7Ge6 crystal was found during the exploration of the electron-poor Gd5−xEuxGe4 phases with x > 3.14 The actual composition for the crystal could not be resolved by the conventional X-ray/neutron diffraction techniques, because of the neighboring atomic numbers and large neutron cross sections of Gd and Eu.15 Following the discovery of the R7Ge6 crystal, the Eu7−xGdxGe6 (x = 0−2) samples were prepared. The starting materials were pieces of europium (99.9+ wt %, distilled grade, Rhone Poulenc), gadolinium (99.99 wt %, distilled grade, Metal Rare Earth Limited, China), and germanium (99.9999 wt %, Alfa Aesar). Because of the high air sensitivity of europium metal, all materials were handled in an argon-filled glovebox. The surface of europium metal lumps was scraped with a file before being cut into pieces. Mixtures of each reactant based on the Eu7−xGdxGe6 stoichiometry were loaded into tantalum ampules, which were sealed by arc melting under an argon atmosphere. These ampules were heated in a high-frequency induction furnace at 1450 °C for 4 h and thereafter were subjected to 1200 °C for 15 h. Received: December 7, 2011 Published: February 15, 2012 3172

dx.doi.org/10.1021/ic2026413 | Inorg. Chem. 2012, 51, 3172−3178

Inorganic Chemistry

Article

Table 1. Phase Analysis of Eu7−xGdxGe6 (x = 0−2.0) from the Rietveld Refinement of the X-ray Powder Diffraction Dataa phase type

a (Å)

b (Å)

c (Å)

V (Å3)

phase fraction (wt %)

0

CrB (Cmcm) Cr5B3 (I4/mcm)

4.72971(4) 7.9895(9)

11.2083(1) 7.9895(9)

4.11203(4) 15.397(3)

217.987(4) 982.8(2)

76(1) 24.0(6)

0.25

Ca7Sn6 (Pnma) Gd5Si4 (Pnma) CrB (Cmcm)

7.6152(5) 7.737(2) 4.725(2)

22.993(2) 15.291(4) 11.215(4)

8.3854(6) 8.020(2) 4.111(1)

0.50

Ca7Sn6 (Pnma) Cr5B3 (I4/mcm)

7.615(1) 7.959(8)

22.999(3) 7.959(8)

0.75

Gd5Si4 (Pnma) CrB (Cmcm)

7.7129(9) 4.7149(3)

15.864(2) 11.2011(8)

8.148(1) 4.1048(3)

997.0(2) 216.78(2)

58.5(5) 41.5(3)

1.0

Gd5Si4 (Pnma) CrB (Cmcm)

7.6932(8) 4.7170(4)

15.742(2) 11.2060(9)

8.149(1) 4.1070(3)

986.9(2) 217.09(3)

54.2(5) 45.8(2)

1.3

Gd5Si4 (Pnma) CrB (Cmcm)

7.6728(8) 4.7046(5)

15.615(2) 11.188(1)

8.1164(9) 4.0997(4)

972.4(2) 215.80(3)

54.6(5) 45.4(3)

2.0

Gd5Si4 (Pnma) CrB (Cmcm)

7.642(1) 4.7136(6)

15.492(2) 11.201(1)

8.097(1) 4.1040(6)

958.6(2) 216.68(5)

74.4(6) 25.6(3)

x

a

8.381(1) 15.208(7)

1468.3(2) 948.9(4) 217.8(1)

90.6(7) 8.1(3) 1.2(5)

1467.8(4) 963(1)

92.2(8) 7.8(2)

The Ca7Sn6-type (Eu,Gd)7Ge6, Gd5Si4-type (Eu,Gd)5Ge4, CrB-type (Eu,Gd)Ge, and Cr5B3-type (Eu,Gd)5Ge3 compounds were identified.

X-ray Analysis. The phase purity of the samples was evaluated through powder X-ray diffraction (XRD) on a PANalytical X’Pert Pro diffractometer, using Co Kα radiation. All powder XRD patterns were analyzed through Rietveld refinement, using the Rietica package.16 A pseudo-Voigt function with the Howard asymmetry correction was used to model the peak profile, and the background was described with a polynomial function. The refinement results are gathered in Table 1. The sample with x = 0 was revealed to contain the CrB-type (Cmcm) (Eu,Gd)Ge compound as a major phase and the Cr5B3-type (I4/mcm) (Eu,Gd)5Ge3 compound as a minor phase. The samples with x = 0.25−0.50 yield the dominant Ca7Sn6-type (Eu,Gd)7Ge6 phase. A mixture of the Gd5Si4-type (Eu,Gd)5Ge4 phase with the CrB-type phase was found in the samples with x = 0.75−2.0. Single-crystal XRD data at room temperature were obtained on a STOE IPDS II diffractometer with Mo Kα radiation. A numerical absorption correction was based on the crystal shape obtained from the optical face indexing and then performed by the program X-Shape, which was included with the STOE IPDS software package.17 The space group was assigned on the basis of the systematic absences and the statistical analysis of the intensity distributions. Structure determination using direct methods and refinement using full-matrix least-squares on F2 were carried out with the SHELXL package.18,19 Because of the similar X-ray scattering factor of Gd and Eu, attempts to fix Gd on a specified R-site had little or no improvement on the refinements. Considering that R/R′ mixtures are present on all metal sites, despite of site preferences in the (R,R′)5X4 phases,20 we believe that a similar mixing occurs in R7Ge6 and, thus, the same Gd/Eu statistical mixtures consistent with the loaded compositions are assumed on all of the metal sites. Some details of the single-crystal data collections and refinements are listed in Table 2. Atomic position parameters and isotropic displacement parameters are given in Table 3, and the selected bond distances are gathered in Table 4. Further details of the crystal structure investigated are available from the Fachinformationszentrum Karlsruhe, 76344 Eggenstein-Leopoldshafen, Germany, on quoting the depository number CSD-423666 (Eu6.5Gd0.5Ge6). Magnetic Measurements. Magnetization in a zero-field-cooled mode for the polycrystalline Eu6.5Gd0.5Ge6 sample was measured on a Quantum Design SQUID magnetometer in the temperature range of 2−350 K under a 100 Oe magnetic field. The Curie temperature (TC) was determined as a maximum on the dM/dT curve. The Weiss temperature (θ) and the effective magnetic moment (μR) of R (where

Table 2. Crystallographic Data for Eu6.5Gd0.5Ge6 as Obtained from the Single-Crystal X-ray Diffraction at Room Temperaturea

a b

parameter

value

empirical formulab structure type a b c volume, V Z θ range number of data/parameters goodness-of-fit on F2 R1/wR2 [I > 2σ(I)] extinction coefficient peak hole

Eu6.5Gd0.5Ge6 Ca7Sn6 (Pnma, No. 62) 7.5943(5) Å 22.905(1) Å 8.3610(4) Å 1454.4(1) Å3 4 2.59°−34.65° 3076/65 0.783 0.0417/0.0269 0.000073(6) 2.240 e/Å3 −2.704 e/Å3

Using Mo Kα radiation and a STOE IPDS II diffractometer. The formula is assigned according to the loaded composition.

R is a mixture of Eu and Gd) were obtained by fitting the paramagnetic data to the Curie−Weiss law. The isothermal magnetizations at various temperatures were measured in the fields of up to 5 T. Electronic Structure Calculations. In order to understand the Rsite occupation and structural stability, the electronic structures of the hypothetical “Eu6GdGe6” and “Eu7Ge6” compounds were calculated with a tight-binding, linear-muffin-tin orbital method with the atomic sphere approximation (TB−LMTO−ASA),21 using the Stuttgart program.22 The structural parameters of Eu6GdGe6 and Eu7Ge6 were taken from the structural data of Eu6.5Gd0.5Ge6, which are shown in Tables 2 and 3. For the calculations of Eu6GdGe6, the symmetry was reduced to Pn21a to allow Gd to occupy one-half of either the R1, R2, or R4 sites. The R1 (8d) site was split into two sites, R1A (4a) and R1B (4a), and so were the R2 and R4 sites in the Pn21a symmetry. Gd was placed either into the R1A, R2A, R3, or R4A sites, and the resulting models were designated as 1, 2, 3, and 4, respectively. The atomic coordinates are given in Table S1 in the Supporting 3173

dx.doi.org/10.1021/ic2026413 | Inorg. Chem. 2012, 51, 3172−3178

Inorganic Chemistry

Article

half-filled Eu and Gd 4f orbitals were treated as core states. The selfconsistent charge density was obtained using 128 irreducible k-points in the first Brillouin zone. The tetrahedron method was performed to calculate the electronic density of states (DOS) and crystal orbital Hamilton population (COHP).25

Table 3. Atomic Coordinates and Isotropic Displacement Parameters (Ueq) for Eu6.5Gd0.5Ge6 as Obtained from the Single-Crystal X-ray Diffraction atom a

R1 R2a R3a R4a Ge1 Ge2 Ge3 Ge4

site

x

y

z

Ueq (Å2)

8d 8d 4c 8d 8d 4c 4c 8d

0.14127(4) 0.02200(4) 0.30874(6) 0.33987(4) 0.19269(9) 0.1844(1) 0.4126(1) 0.00365(9)

0.14086(2) 0.66231(2) 0.2500 0.51891(2) 0.59769(4) 0.2500 0.2500 0.05342(4)

0.31835(6) 0.17991(5) 0.02366(7) 0.17654(5) 0.4695(1) 0.6032(2) 0.3842(1) 0.0391(1)

0.01406(9) 0.01016(8) 0.0150(1) 0.01161(8) 0.0123(2) 0.0169(3) 0.0147(3) 0.0111(2)



RESULTS AND DISCUSSION Synthesis and Structural Features. The Rietveld refinements of the powder XRD data for Eu7−xGdxGe6 (x = 0−2.0) showed that two samples with x = 0.25 and 0.50 had the major phase isostructural with Ca7Sn6 (Table 1). No such phase was seen in the pure Eu7Ge6 sample or in the samples with x ≥ 0.75. Considering the noticeably larger ionic radius of Eu2+ (1.17 Å for 6−coordinate environment) than that of Gd3+ (0.938 Å),26 one expects to see a larger unit cell for the sample with x = 0.25 than that for x = 0.50, because of a greater Eu loading amount. However, the Ca7Sn6-type phases in both the samples have almost identical unit-cell volumes1468.3(2) Å3 for x = 0.25 vs 1467.8(4) Å3 for x = 0.50 (Table 1)indicating very close experimental Eu/Gd ratios. Therefore, we can conclude that the Ca7Sn6-type Eu7−xGdxGe6 phase survives in a narrow compositional range at x ≈ 0.50, and we will focus our discussion on the Eu6.5Gd0.5Ge6 phase. Hitherto, the Ca7Sn6type structure has been found in only two binaries, i.e., Ca7Sn627 and Ca7Ge6.28 Eu6.5Gd0.5Ge6 is the first magnetic pseudo-binary to adopt this structure. The Eu6.5Gd0.5Ge6 compound displays a significant structural similarity to the R5X4 layered structures. The structure can be viewed as an intergrowth of the 2∝ [R5Ge4] slabs and 2∝ [RGe] layers along the b-axis (see Figure 1). The 2∝ [R5Ge4] slabs are interconnected by zigzag Ge4 tetramers. The two building blocks share the same planar R 32434 nets (see Figure 2);

a

R is defined as a statistical mixture of europium and gadolinium (Eu:Gd = 6.5:0.5).

Table 4. Interatomic Distances in Eu6.5Gd0.5Ge6

a

atomsa

distance (Å)

atomsa

distance (Å)

Ge1−Ge4 (×8)

2.586(1)

Ge2−Ge3 (×4)

2.521(2)

Ge4−Ge4 (×4)

2.534(1)

Ge1−R1 (×8) R1 (×8) R2 (×8) R2 (×8) R3 (×8) R4 (×8) R4 (×8) R4 (×8)

3.2492(9) 3.328(1) 3.1196(9) 3.1630(8) 3.5180(9) 3.192(1) 3.2410(9) 3.4534(9)

Ge4−R1 (×8) R1 (×8) R2 (×8) R4 (×8) R4 (×8) R4 (×8) R4 (×8)

3.249(1) 3.6063(9) 3.100(1) 3.1534(9) 3.2534(9) 3.2683(8) 3.3506(9)

Ge2−R1 (×8) R2 (×8) R2 (×8) R3 (×4) R3 (×4)

3.468(1) 3.0687(8) 3.127(1) 3.044(1) 3.640(1)

R1−R1 (×8) R2 (×8) R2 (×8) R2 (×8) R3 (×8) R3 (×8) R4 (×8) R4 (×8) R4 (×8)

3.9654(5) 3.9899(6) 4.3742(7) 4.4019(7) 3.7333(6) 3.7910(5) 4.0978(6) 4.1319(6) 4.3166(6)

Ge3−R1 (×8) R1 (×8) R2 (×8) R3 (×4) R3 (×4)

3.2862(8) 3.4830(9) 3.224(1) 3.105(1) 3.116(1)

R2−R2 (×8) R2 (×4) R3 (×8) R3 (×8) R4 (×8) R4 (×8)

3.9739(5) 4.0173(6) 3.6388(6) 3.7346(6) 3.7606(6) 4.0763(5)

R4−R4 (×4) R4 (×8) R4 (×8)

3.9218(6) 3.9909(5) 4.4822(6)

Figure 2. Projections of two building fragments, the 2∝ [R5Ge4] slab and 2∝ [RGe] layer, in R7Ge6 along the b-axis. The two fragments contain the eclipsed 32434 nets formed by the R atoms. (R is a mixture of Eu and Gd.)

therefore, it is not surprising that they can be incorporated into one structure. In the 2∝ [R5Ge4] slab, the two adjacent R 32434 nets are staked in an eclipsed fashion, which creates cubic and trigonal prismatic voids. R3 atoms occupy the cubic holes, while Ge2/Ge3 atoms occupy all trigonal prismatic holes. In the 2∝ [RGe] layer, Ge1/Ge4 atoms cap the R triangle and rectangular centers. The Ge−Ge bonds within the Ge4 tetramer are 2.534(1) and 2.586(1) Å (Table 4), which indicate their strong covalent character and compare well with the data reported for Ge single bonds in EuGe2,29 Gd5Si4-type (R,A)5Ge4 (R = rare earth, A = alkaline earth),6−8 and DyGe3.30 The conventional diffraction techniques could not establish the site occupation by the Eu and Gd atoms. However, as shown for the R5X4 systems,20 analysis employing the geometric and electronic considerations may shed light on the R-site occupation. First, the local coordination environments of the four R-sites of Eu6.5Gd0.5Ge6 are analyzed. The R1

The number of bonds per unit cell is given in parentheses.

Information. Exchange and correlation were treated by the local density approximation.23 All relativistic effects except spin−orbit coupling were taken into account, using a scalar relativistic approximation.24 In the ASA method, space is filled with overlapping Wigner−Seitz (WS) atomic spheres whose radii were obtained by requiring the overlapping potential to be the best possible approximation to the full potential. No empty spheres were needed to achieve convergence. In the four models of Eu6GdGe6, the WS radii of R1A(B), R2A(B), R3, R4A(B), Ge1, Ge2, Ge3, and Ge4 were fixed at 2.21, 2.07, 2.05, 2.15, 1.51, 1.45, 1.45, and 1.45 Å, respectively. The WS radii employed for the Eu7Ge6 calculations are as follows: Eu = 2.04−2.21 Å and Ge = 1.45−1.51 Å. The basis sets included 6s, 6p, and 5d orbitals for Eu/Gd, and 4s, 4p, and 4d orbitals for Ge. The Eu/Gd 6p and Ge 4d orbitals were treated by the Löwdin downfolding technique.21,22 The Eu atom was treated as a divalent cation, which is supported by the structural analysis and magnetic measurements. The 3174

dx.doi.org/10.1021/ic2026413 | Inorg. Chem. 2012, 51, 3172−3178

Inorganic Chemistry

Article

Table 5. Total Electronic Energy (Etot, eV/cell) and the Integration of Crystal Orbital Hamilton Population (−ICOHP) for Various Bonds in the Structural Models of Eu6GdGe6a

site is surrounded by 10 R and 7 Ge atoms; the R2 site is coordinated by 10 R and 6 Ge atoms; the coordination polyhedron around R3 can be seen as a combination of a distorted R cube and a Ge octahedron; and the R4 site is surrounded by 10 R and 7 Ge atoms (see Figure 3). One can

bond strength (eV/cell) Gd−Ge Gd−R Eu−Ge Eu−Eu Ge−Ge Sum(−ICOHP) Etot (eV/cell)

Model 1

Model 2

Model 3

Model 4

6.80 2.54 33.71 5.57 10.04 58.66 0.21

7.67 2.69 33.17 5.40 10.00 58.93 0

6.62 2.44 34.04 5.48 10.07 58.65 0.16

7.09 2.53 33.50 5.65 10.01 58.78 0.10

a

The lowest-energy terms are indicated in boldface font within the table.

total electronic energy (Etot); the Etot values for models 1, 3, and 4 are, respectively, 0.21, 0.16, and 0.10 eV/cell higher than that for model 2. Therefore, the Gd atoms are predicted to occupy the slab-surface R2 site, based on the geometric and electronic considerations. Magnetic Properties. To understand the structural stability and bonding character of Eu6.5Gd0.5Ge6, one must establish the oxidation state of Eu. Eu6.5Gd0.5Ge6 (V = 1454.4(1) Å3) displays a larger unit cell than Ca7Ge6 (V = 1339.2(3) Å3),28 implying a divalent Eu species, according to the Shannon radius order of Eu2+ (1.17 Å) > Ca2+ (1.00 Å) > Eu3+ (0.947 Å) > Gd3+ (0.938 Å).26 This oxidation state for Eu is also verified through the magnetic data. A fit to the paramagnetic data using the Curie−Weiss law yielded an effective magnetic moment (μeff) of 8.24 μB for Eu, assuming Gd takes the theoretical moment of 7.94 μB (see Figure 4).

Figure 3. Coordination polyhedral around the four rare-earth sites of Eu6.5Gd0.5Ge6. The site volume of the Wigner−Seitz cells is indicated.

see that the coordination polyhedral around the R1, R2, and R3 sites exhibit considerable similarities to the corresponding R-sites of R5X4,20 because these three R-sites belong to the same building unit, the 2∝ [R5X4] slab, in both structures. A subtle difference between these two structures lies in the site size, i.e., the slab-centered R3 site of R5X4 has the tightest environment, whereas the slab-surface R2 site of Eu6.5Gd0.5Ge6 has the smallest site volume (28.75 Å3), as calculated by DIDO95.31 The smaller Gd atom, as compared to Eu (the Shannon radius order of Gd3+ (0.938 Å) < Eu3+ (0.947 Å) < Eu2+ (1.17 Å) for six-coordinate environments26), is expected to prefer the smallest R2 site. The average R−Ge bond lengths are, respectively, 3.381(1) Å, 3.134(1) Å, 3.324(1) Å, and 3.273(1) Å for the R1, R2, R3, and R4 sites, while these sites have the average R−R distances of 4.0767(6), 3.9728(7), 3.7245(6), and 4.1251(6) Å, respectively. Obviously, the R2 site has the shortest R−Ge bonds, whereas the R3 site shows the shortest R−R bonds. The metal-site occupation in intermetallics can also be considered from the perspective of electronic effects. The atoms with more electonegativy tend to occupy the sites with the larger Mulliken population.32 In the TB-LMTO-ASA calculations, the Mulliken population can be treated as the integrated DOS (IDOS) in a semiempirical way.9 For the calculations of the hypothetic “Eu7Ge6” phase, the IDOS for the R1, R2, R3, and R4 sites are 2.45, 2.58, 2.08, and 2.48, respectively. Since Gd has a larger electronegativity than Eu (Allred−Rochow electronegativity of χ(Gd) = 1.11 versus χ(Eu) = 1.01),33 the Gd atoms are expected to occupy the R2 site with the largest IDOS value. The Gd-substitution effects on the bonding and electronic structures were examined through the calculations of “Eu6GdGe6”. The bond strength is scaled by means of the integration of crystal orbital Hamilton population (−ICOHP). The assignment of Gd at the R2 site (model 2) yields the strongest Gd−Ge and Gd−R interactions, leading to the lowest bond energy, Sum(−ICOHP) = 58.93 eV/cell, but the Eu−Ge and Eu−Eu interations are slightly weakened in Model 2 (see Table 5). The strong bond interactions around R2 correspond to its short R−Ge and R−R bond distances. Also, the presence of Gd on the R2 site results in the lowest

Figure 4. Temperature (T) dependence of magnetization (M) and inverse magnetic susceptibility (1/χ) of Eu6.5Gd0.5Ge6 in a magnetic field of 100 Oe. The Weiss temperature (θ) and effective magnetic moment of R (μR) are derived from a fit (denoted by a solid line in the figure) to the paramagnetic data, using the Curie−Weiss law. (R is a mixture of Eu and Gd.)

Since μeff(Eu2+) = 7.94 μB and μeff(Eu3+) = 0 μB for free Eu ions, the data unequivocally advocate the presence of Eu2+ in Eu6.5Gd0.5Ge6. The overestimation of μeff(Eu) results from the 5d-electron polarization, which is frequently observed in the rare-earth intermetallics.34,35 The magnetic properties of Eu6.5Gd0.5Ge6 were studied further by measuring its isothermal magnetization at various temperatures (see Figure 5). The magnetization at low temperatures, such as at 2 K, is not saturated up to 5 T and 3175

dx.doi.org/10.1021/ic2026413 | Inorg. Chem. 2012, 51, 3172−3178

Inorganic Chemistry

Article

the measured magnetic moment is ∼2.8 μB/R in 5 T, which is much smaller than the saturation magnetic moment of 7 μB for Eu2+/Gd3+ (1 gJ = 7 μB). Arrott plots below the critical temperature (TC = 45 K) display a distinct convex curvature and an absence of spontaneous magnetization (see Figure 5),

tend to stabilize the structures, irrespective of the changes in the valence electron concentration (VEC).5 However, the Ge− Ge bonds have significant antibonding contribution at the EF level and are prone to cleavage, depending on the VEC.5 We start our analysis with a qualitative approach based on the Zintl−Klemm concept with a focus on the Ge−Ge interactions.41−43 Within this formalism, the intraslab Ge2− Ge3 dimers, located within 2∝ [R5Ge4] slabs are isoelectronic with halogen ones and carry a negative charge of −6. The interslab Ge1−Ge4−Ge4−Ge1 tetramer can be treated as a fragment cut from a one-dimensional 1∝ [Ge2−] chain observed in EuGe and DyGe3,30,43 and it will have a configuration of Ge13−−Ge42−−Ge42−−Ge13−, with a total charge of 10−. Thus, the electronic formula of Eu6.5Gd0.5Ge6 can be written as (Eu2+)6.5(Gd3+)0.5(Ge26−)(Ge410−)(1.5 h+), suggesting presence of 1.5 holes in the valence band. Obviously, this approach is simplified, because it ignores R−R and R−Ge interactions, both of which will be significant and bonding up to the EF level (see Figure 6). Still, the electron deficiency derived

Figure 5. Magnetization (M) vs magnetic field (H) (top panel) and Arrott plots (bottom panel) of Eu6.5Gd0.5Ge6 at various temperatures.

indicating a disordered ferromagnetic (FM) state in Eu6.5Gd0.5Ge6.36 It could be helimagnetic, i.e., the magnetization of the 2∝ [R5Ge4] slabs is spirally coupled to that of the adjacent2∝ [RGe] layers, which was observed in some rare-earth metals and compounds with layered structures, e.g., Dy,37 Tb5Ge3,38 Ce2(Fe,Mn)17.39 However, further investigations are needed to fully understand the magnetic structure of this material. At low temperatures (T < 20 K), a metamagnetic transition occurs in high fields, e.g., at ∼40 kOe at 2 K and ∼34 kOe at 17 K (see Figure 5). The positive slopes of the Arrott plots before and after the metamagnetic transition, as well as around TC (see Figure 5), suggests that the magnetic transitions are second order.40 Electronic Structure and Bonding. One of the interesting points is what dictates the stability and composition of the Eu6.5Gd0.5Ge6 phase. The 2∝ [R5Ge4] slabs and 2∝ [RGe] layers are found in the corresponding pseudo-binaries, such as (Gd,Eu)5Ge4 and (Gd,Eu)Ge (Eu5Ge4 is not known, but can be stabilized through Gd substitution).14 Obviously, when these building blocks are fused, the resulting new bonds are likely to dictate the phase stability and composition, although changes to the existing interactions also can be a factor. The interactions in Eu6.5Gd0.5Ge6 can be divided into three groups: R−R, R−Ge, and Ge−Ge. Studies on R5Ge4 and RGe showed that R−R and R−Ge interactions are bonding at the Fermi level (EF) and

Figure 6. Density of states (DOS) and crystal orbital Hamilton population (COHP) curves for the hypothetical “Eu7Ge6” compound. The Ge2 and Ge3 atoms have similar projected DOS; Ge1 and Ge4 also have similar projected DOS. The vertical line at 0.07 eV is the Fermi level (EF) for Eu7−xGdxGe6 (x = 0.5) within a rigid band approximation. The dashed line, locating at a pseudo-gap, denotes the states containing 40 valence electrons per formula unit. Interactions with −COHP > 0 and −COHP < 0 are bonding and antibonding, respectively.

by this method is reflected in the electronic structure (see Figure 6), calculated using the TB−LMTO−ASA program. A pseudo-gap at 0.61 eV corresponds to 40 e−/formula, which is associated with a hypothetical closed-shell (Eu2+)5(Gd3+)2(Ge6)16− phase with the Ca7Sn6 structure. The 3176

dx.doi.org/10.1021/ic2026413 | Inorg. Chem. 2012, 51, 3172−3178

Inorganic Chemistry

Article

closed-shell configuration of “Eu5Gd2Ge6” is expected to generate a band gap. However, the oversimplified Zintl− Klemm formalism ignores the existence of R 5d orbitals and their dispersion characteristics around the EF (see Figure 6). The significant hybridization among the R d-states and the Ge p-states leads to the obvious R−R and R−Ge bonding interactions around the EF (Figure 6), eliminating the expected gap and resulting instead in a pseudo-gap for “Eu5Gd2Ge6”. The contribution from the metal d orbitals and their bonding with the neighboring main-group-element p orbitals around EF were used to explain metallic properties of the nominally electron-precise Ca5Ge3 compound and related Cr5B3-type phases.45 The Fermi level of Eu6.5Gd0.5Ge6 is observed at 0.07 eV (x = 0.5) and lies below the pseudo-gap, confirming its electron-poor character. The questions that can be asked are: (i) What is the effect of this lower VEC on bonding? and (ii) Why does its phase range seem to be limited? For the Ge26− dimer of the electron-precise R7Ge6 phase, the molecular orbitals (MOs) can be simply written as σs2σs*2σp2πp4πp*4σp*0. The πp and πp* orbitals have a relatively small contribution to the dimer bonding, because the π interactions are relatively weak and the p orbitals contributing to these π MOs are strongly involved into the R−Ge interactions. Still, the πp* states of the Ge2−Ge3 dimer are clearly seen in the COHP plot (see Figure 6). It is worth mentioning that, since the Ge2−Ge3 dimer resides within the 2∝ [R5Ge4] slab, which is a building block for the R5X4 phases; its bonding character is similar to that of the X−X dimers in R5X4.5 However, a different electron count yields different population of the p-states for the dimers: in Eu6.5Gd0.5Ge6, EF falls into the πp* states, whereas in R5X4, e.g., Gd5Ge4, EF is pushed up into the σp* states, because of an electron excess. The bonding within the Ge4 tetramer can be also analyzed in terms of individual Ge1−Ge4 and Ge4−Ge4 interactions (see Figure 6). However, there is a striking difference, especially in the area of the σs states, between these dimers and Ge2−Ge3. Namely, there are four σs peaks for the Ge1−Ge4 and Ge4− Ge4 bonds and only two σs peaks for the Ge2−Ge3 dimer. This difference is due to the fact that the Ge1−Ge4 and Ge4−Ge4 dimers are fused into a tetramer with the local C2 h symmetry. The resulting MOs for the σs and πp interactions are shown in Figure 7. Interestingly, the σs, σs*, πp, and πp* states will have different sequences for the Ge1−Ge4 and Ge4−Ge4 dimers: bonding, bonding, antibonding, and antibonding for the Ge1− Ge4 dimers and bonding, antibonding, bonding and antibonding for the Ge4−Ge4 ones. These trends are clearly seen in the COHP plots in the range from −10 eV to −6 eV, where the s states dominate (see Figure 6). As for the π states, these features are washed out due to the weaker π interactions and also due to a significant mixing with the d orbitals of neighboring R atoms. Still, an antibonding character of the Ge1−Ge4 interactions above −0.8 eV is visible and stems from the πp* states. Thus, the Ge−Ge bonds would promote a lower VEC, which would minimize the population of the πp* states. However, a VEC reduction would have somewhat different effects on different bonds. A lower VEC of 38 e−, as in pure Eu7Ge6, will strengthen the Ge1−Ge4 interactions within the tetramer but will have almost no effect on the Ge4−Ge4 within the tetramers. In fact, it appears that the overall bonding within the tetramer can tolerate even a slightly higher VEC as in Eu6.5Gd0.5Ge6 without much penalty. A lower VEC will benefit

Figure 7. Schematic molecular orbital diagram for 4s and 4p π electrons of a Ge2 dimer (Ge2−Ge3) and zigzag Ge4 tetramer (Ge1− Ge4−Ge4−Ge1).

most the Ge2−Ge3 dimer as fewer of its strongly antibonding πp* states will be populated. On the other hand, any decrease in the VEC, achieved through a lower Gd amount or Gd absence, will weaken the bonding R−Ge and R−R. In addition, the Gdfree Eu7Ge6 phase is compositionally close to the EuGe binary, which contains infinite Ge zigzag chains with the Ge−Ge bonding resembling that of the Ge tetramers in Eu6.5Gd0.5Ge6. Thus, a larger VEC achieved through Gd substitution is required to stabilize the Ca7Sn6-type structure. On the other hand, an even larger VEC (beyond x = 0.5 of Gd) appears to promote formation of the Gd5Si4-type R5Ge4 phases, which offer more of the strong R−R bonds in the presence of weaker Ge−Ge dimers. In summary, one may argue that there is a fine balance between the VEC and bonding in Eu6.5Gd0.5Ge6 and the VEC of 38.5 e− is optimal for the overall bonding.



CONCLUSIONS The electron-poor Eu6.5Gd0.5Ge6 polar intermetallics phase adopts the orthorhombic Ca7Sn6-type structure and it can be considered as an intergrowth of the Gd5Si4-type (Pnma) R5Ge4 and FeB-type (Pnma) RGe phases. The slab-surface R2 site is expected to preferentially accommodate Gd, based on the geometric and electronic factors. The Ca7Sn6-type (Eu,Gd)7Ge6 phase survives in a narrow composition range. Its stability and composition represents a compromise between the VEC, bonding, and competition from the binary EuGe and pseudobinary (Eu,Gd)5Ge 4 phases. Eu6.5Gd0.5Ge 6 provides an attractive way to develop new compounds by integrating the 2∝ [R5X4] slabs with other symmetry-compatible building blocks. Furthermore, considering the structural similarity of the 3177

dx.doi.org/10.1021/ic2026413 | Inorg. Chem. 2012, 51, 3172−3178

Inorganic Chemistry

Article

(24) Andersen, O. K.; Jepsen, O.; Glotzel, D. In Highlights of Condensed-Matter Theory; Bassani, F., Fumi, F., Tosi, M. P., Eds.; Elsevier: Amsterdam, 1985; p 59. (25) Blöchl, P. E.; Jepsen, O.; Andersen, O. K. Phys. Rev. B 1994, 49, 16223. (26) Shannon, R. D. Acta Crystallogr., Sect. A: Cryst. Phys. Diffr. Theor. Gen. Crystallogr. 1976, 32, 751. (27) Palenzona, A.; Manfrinetti, P.; Fornasini, M. L. J. Alloys Compd. 2000, 312, 165. (28) Palenzona, A.; Manfrinetti, P.; Fornasini, M. L. J. Alloys Compd. 2002, 345, 144. (29) Bobev, S.; Bauer, E. D.; Thompson, J. D.; Sarrao, J. L.; Miller, G. J.; Eck, B.; Dronskowski, R. J. Solid State Chem. 2004, 177, 3545. (30) Schobingerpapamantellos, P.; Demooij, D. B.; Buschow, K. H. J. J. Alloys Compd. 1992, 183, 181. (31) Koch, E.; Fischer, W. Z. Kristallogr. 1996, 211, 251. (32) Gordon, J. M. Eur. J. Inorg. Chem. 1998, 1998, 523. (33) James, A. M.; Lord, M. P. Macmillan’s Chemical and Physical Data; Macmillan: London, UK, 1992. (34) Harmon, B. N.; Freeman, A. J. Phys. Rev. B 1974, 10, 1979. (35) Roeland, L. W.; Cock, G. J.; Muller, F. A.; Moleman, A. C.; McEwen, K. A.; Jordan, R. G.; Jones, D. W. J. Phys. F 1975, 5, L233. (36) Yeung, I.; Roshko, R. M.; Williams, G. Phys. Rev. B 1986, 34, 3456. (37) Behrendt, D. R.; Legvold, S.; Spedding, F. H. Phys. Rev. 1958, 109, 1544. (38) Schobinger-Papamantellos, P. J. Magn. Magn. Mater. 1982, 28, 97. (39) Prokhnenko, O.; Arnold, Z.; Kamarad, J.; Ritter, C.; Isnard, O.; Kuchin, A. J. Appl. Phys. 2005, 97, 113909. (40) Banerjee, B. K. Phys. Lett. 1964, 12, 16. (41) Zintl, E. Angew. Chem. 1939, 52, 1. (42) Schäfer, H.; Eisenmann, B.; Müller, W. Angew. Chem., Int. Ed. 1973, 12, 694. (43) Chemistry, Structure, and Bonding of Zintl Phases and Ions; Kauzlarich, S. M. ed.; VCH Publishers: New York, 1996. (44) Albert, K.; Meyer, H. J.; Hoffmann, R. J. Solid State Chem. 1993, 106, 201. (45) Mudring, A. V.; Corbett, J. D. J. Am. Chem. Soc. 2004, 126, 5277.

R5X4 and R7X6 structures, we may anticipate structural transformations and concomitant changes in the physical properties via controlling the cleavage/formation of the X−X dimers.



ASSOCIATED CONTENT

S Supporting Information *

Single-crystal crystallographic files in CIF format, and the structural models for Eu6GdGe6. This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Author

*Tel.: +1 905 525 9140, ext 27796. Fax: +1 905 521 2773. E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by Discovery Grant from the Natural Sciences and Engineering Research Council of Canada and by a grant from the ACS Petroleum Research Fund.



REFERENCES

(1) Pecharsky, V. K.; Gschneidner, K. A. Phys. Rev. Lett. 1997, 78, 4494. (2) Choe, W.; Pecharsky, V. K.; Pecharsky, A. O.; Gschneidner, K. A.; Young, V. G.; Miller, G. J. Phys. Rev. Lett. 2000, 84, 4617. (3) Pecharsky, V. K.; Holm, A. P.; Gschneidner, K. A.; Rink, R. Phys. Rev. Lett. 2003, 91, 197204. (4) Magen, C.; Arnold, Z.; Morellon, L.; Skorokhod, Y.; Algarabel, P. A.; Ibarra, M. R.; Kamarad, J. Phys. Rev. Lett. 2003, 91, 207202. (5) Mozharivskyj, Y.; Choe, W.; Pecharsky, A. O.; Miller, G. J. J. Am. Chem. Soc. 2003, 125, 15183. (6) Wu, L. M.; Kim, S. H.; Seo, D. K. J. Am. Chem. Soc. 2005, 127, 15682. (7) Tobash, P. H.; Bobev, S. J. Am. Chem. Soc. 2006, 128, 3532. (8) Tobash, P. H.; Bobev, S.; Thompson, J. D.; Sarrao, J. L. Inorg. Chem. 2009, 48, 6641. (9) Misra, S.; Miller, G. J. J. Am. Chem. Soc. 2008, 130, 13900. (10) Svitlyk, V.; Campbell, B. J.; Mozharivskyj, Y. Inorg. Chem. 2009, 48, 10364. (11) Kozlov, A. Y.; Pavlyuk, V. V.; Davydov, V. M. Intermetallics 2004, 12, 151. (12) Holtzberg, F.; Gambino, R. J.; McGuire, T. R. J. Phys. Chem. Solids 1967, 28, 2283. (13) Smith, G. S.; Tharp, A. G.; Johnson, W. Acta Crystallogr. 1967, 22, 940. (14) Yao, J.; Wang, P.; Mozharivskyj, Y. Chem. Mater. 2012, DOI: 10.1021/cm203148e. (15) Sears, V. F. Neutron News 1992, 3, 26. (16) Hunter, B. A. Int. Union Crystallogr. Comm. Powder Diffr. Newsl. 1998, 20, 21. (17) STOE; 2.05 ed.; STOE & Cie GmbH: Darmstadt, Germany, 2004. (18) Sheldrick, G. M. Acta Crystallogr., Sect. A: Found Crystallogr. 2008, 64, 112. (19) Sheldrick, G. M. SHELXL; University of Göttingen: Göttingen, Germany, 1997. (20) Yao, J.; Mozharivskyj, Y. Z. Anorg. Allg. Chem. 2011, 637, 2039. (21) Lambrecht, W. R. L.; Andersen, O. K. Phys. Rev. B 1986, 34, 2439. (22) Jepsen, O.; Burkhardt, A.; Andersen, O. K.; The TB-LMTOASA Program, 4.7 ed.; Max-Planck-Institut fr Festkrperforschung: Stuttgart, Germany, 1999. (23) Andersen, O. K.; Jepsen, O. Phys. Rev. Lett. 1984, 53, 2571. 3178

dx.doi.org/10.1021/ic2026413 | Inorg. Chem. 2012, 51, 3172−3178